You are on page 1of 10

Alpha particles

Gold foil

Radium Lead collimator

Zinc sulfide screen

Zinc sulfide screen

Rutherford Scattering
1 Introduction

It was quite impossible to explain the large-angle scattering of an alpha particle in terms of a large number of small-angle scatterings. Geiger and Marsden found in 1909 that the most probable angle by which alpha particles from radium C (the granddaughter of natural radium) are scattered in passing through a thin gold foil (4105 cm thick) is 0.87o, but about one in every 20,000 alpha particles is scattered backward (that is, by more than 90o , which is more than 100 times the most probably angle). A well-known theorem in the mathematical theory of probability, the central limit theorem, gives a formula for the probability of nding any specic value for a quantity that is made up of many statistically independent small increments, each of which can be in any direction. According to this formula, the probability of nding such a quantity to have a value more than 100 times its most probably value (or, strictly speaking, more than 100 times its root-mean-square value) is only 3102174 . Even if all the material of the universe consisted of alpha particles and each alpha particle was red billions of times per sec1

ond through this gold foil, the chance that such an improbably event would have occurred even once in the history of the universe would still be utterly negligible. As Rutherford concluded, these large-angle scatterings could only be explained if there is an appreciable probability that an alpha particle is deected by a large angle in a single encounter with an atom. The important thing about Rutherfords work is not just that he had gotten the right idea - that an atom consists of a small, heavy, positively charged nucleus surrounded by orbiting electrons - but that he had found a way to test it. The analysis Rutherford used is one that has been repeated countless times since 1911 in studies of the structure of atoms, nuclei, and elementary particles. Suppose we want to test some hypothesis about the nature of the atom, such as Rutherfords picture of a tiny positive nucleus surrounded by a cloud of electrons. Using this hypothesis together with Newtonian mechanics, we can calculate the hyperbolic orbit of an alpha particle red at the atom, in much the same way that an astronomer calculates the hyperbolic orbit of a comet passing through the solar system. Of course, one cannot see into the atom, so the interesting thing is the one thing that can be measured: the scattering angle, the angle between the initial direction of the alpha particle as it comes in from innity and the direction along which it recedes to innity again after the encounter. but unfortunately this scattering angle is not xed; it depends on the line along which the alpha particle approaches the atom. It is convenient to express this dependence in terms of the impact parameter, the distance by which the alpha particle would miss the center of the nucleus if it were not deected. For instance, for an alpha particle with velocity 2.09107 meters per second approaching a neucleus with an electric charge of Z electronic charges, the scattering angle for an impact parameter of 1.51016 meters can be calculated to be 90o. (See section ?? for the formula for carrying out such calculations. Incidentally, it is no coincidence that the impact parameter that gives a large scattering angle like 90o is of the same order of magnitude as the distance of closest approach for a head-on encounter. In both of these situations the alpha particle gets close enough to the nucleus that its initial kinetic energy is largely used up in doing work against the electrical repulsion of the nucleus, a necessary condition if it is to be strongly deected.) How can we possibly use the results of such calculations to analyze experimental data? After all, the alpha particles are not aimed at specic atoms, but are just red blindly into a foil containing a vast number of invisible 2

atoms. The answer found by Rutherford is that the analysis must be done statistically, not by measuring the scattering angle for a single alpha particle encountering an atom at a known impact parameter, but by measuring the distribution of scattering angles for many alpha particles that happen to pass close to one atom or another at random impact parameters. For instance, suppose we measure the fraction of all alpha particles scattered by at least a given angle, say 1o or 90o or 179o or whatever. In order for this to happen, the impact parameter would have to be less than a certain amount; in the example above, it would have to be less than 1.5Z1016 metters for tha alpha particle to be scattered by at least 90o . For the purpose of calculating the fraction of alpha particles scattered by at least a given angle, each nucleus can be thought of as a little disc facing the incoming alpha particle, the disc radius being the maximum impact parameter for such scattering: Only those alpha particles that happen to hit one of these discs will be scattered by at least the given angle. The fraction of alpha particles scattered by at least the given angle is thus simply equal to the fraction of the area of the foil occupied by these discs - in other words, by the area of each disc times the average number of atoms per unit area in the foil. By the familiar formula for the area of a circle, the area of each disc is times the square of the maximum impact parameter for scattering by at least the given angle. This area depends on the scattering angle we are interested in. It is clearly not the actual area of any physical disc, but it is the fundamental quantity that determines the probabilities of scattering by various angles, and is therefore called the eective cross section of the atom. A good deal of modern physics consists of the measurement of such cross sections. For example, we saw that the maximum impact parameter of scattering of an alpha particle by at least 90o in the Geiger-Marsden experiments was calculated to be 1.5Z1016 meters. (Recall that Z is the charge of the nucleus in units of the charge of the electron.) The eective cross section was therefore (1.5 Z 1016 m)2 = 7 Z 2 1032 m2 Also, the number of gold atoms in a square meter of foil is calculated by taking the mass per square meter of foil, which is the density 1.93104 kg/m3 of gold times the thickness 4107 m of the foil, and dividing this by the mass of one gold atom, which is the atomic weight 197 of gold times the mass

1.71027 kg of a unit atomic weight. This gives (1.93 104 kg/m3 ) (4 107 m) = 2.3 1022 atoms/m2 197 (1.7 1027 kg) In one square meter of foil the total area occupied by our ctitious little discs is then the number 2.31022 of atoms times the area 7Z2 1032 m2 of each disc, or 1.6109 Z2 square meters, so the probability that the alpha particle happens to be aimed at one of these ctitious discs and will therefore be scattered by at least 90o is 1.6109 Z2 . (The fact that this is much less than 1 shows that we can ignore the possibility of some discs overlapping.) In comparison, Geiger and Marsden had measured this probability to be about one in 20,000, or 5105 , so it was possible to conclude that Z, the electric charge of the nucleus, must be approximately Z 5 105 = 180 1.6 109

This is not too good a value; we now know that the nucleus of the gold atom has an electric charge of 79 electronic units. However, Geiger and Marsden had not in 1909 aimed at a precise measurement of scattering probabilities, so the discrepancy is not surprising. Rutherford in his 1911 paper actually used somewhat more precise data of Geiger and Marsden on small-angle scattering of alpha particles, and found values for the nuclear charge Z of gold of 97 in one case and 114 in another. He also used data of J. A. Crowther on the scattering of beta rays to determine Z for various other elecments. I do not know why Rutherfords results for Z were systematically too high, but at least they were of the right order of magnitude, and showed that nuclear charge increases with atomic weight, as might have been expected. Much more important than these crude measurements of nuclear charge was the verication of Rutherfords basic assumption that the scattering is due to a small, heavy, charged nucleus. Rutherford had calculated the impact parameter that would give scattering by a given angle; squaring this and multiplying by then gave the eective cross section for scattering by that angle or more.1 For instance, according to Rutherfords formula, the eective
Rutherford was lucky to get the right answer for the relation between impact parameter and scattering angle. In general such calculations have to be carried out by the methods of quantum mechanics, and for the energies and masses typical of nuclear physics the results are very dierent from those that would be obtained by Rutherfords approach, in which one calculates the orbits of the scattered particles by the rules of classical Neutonian mechanics. It happens that there is just one case for which the quantum and the classical
1

cross section for scattering by at least 135o is less than that for scattering by at least 90o by a factor 0.00196. As we have seen, the fraction of alpha particles scattered by various angles are just given by the product of these cross sections times the number of atoms per unit area of foil. Starting in 1911, Geiger and Marsden began a program of carefully measuring the fraction of alpha particles scattered by various angles, and in 1913 they reported that their experimental results were in good agreement with Rutherfords theoretical formulas. Thus, the correctness of Rutherfords picture of an atomic nucleus surrounded by electrons was now denitely established.

Calculate deection angle

Here we will describe the formula derived by Rutherford for the scattering of an alpha particle by an atomic nucleus, and will show how this formula was used to verify the existence of the nucleus and to measure its charge. Suppose an alpha particle is red at an atom, in such a way that if it were not deected it would miss the nucleus by a distance b. The quantity, the distance of closest approach if the forces between the alpha particle and the nucleus were miracuously turned o, is known as the impact parameter. By applying Newtons Second Law to the motion of the alpha particle, for each value of the impact parameter we can calculate the scattering angle , the angle between the initial and the nal directions of the alpha particles velocity, see Figure 1. We will not be able here to go through the details of this calculation, but fortunately we can go a long way toward the nal answer by a method of reasoning known as dimensional analysis. This method is based on the principle that the value of whatever quantity we are trying to calculate cannot depend on the units used in measuring the other quantities on which it depends. Rutherford scattering provides a nice example of the power and limitations of this method. First, we must consider what are the input parameters on which the scattering angle might depend. It will certainly depend on the impact parameter
approaches give precisely the same answer for scattering problems: the case of forces that decrease as the inverse square of the distance, which of course is just the case that was of interest to Rutherford. If Thomsons plum pudding model of the atom had been correct, then classical calculations like those of Rutherford would have given the wrong answer for the cross section, and it would have been impossible to interpret the results of the Geiger-Marsden experiment correctly until the development of quantum mechanics.

Outgoing particle

Incoming particle

b
Target particle

Figure 1: Schematic diagram of a scattering event, showing the denition of impact parameter (b) and deection angle ().

b, and on the initial speed v of the alpha particle. Also, by combining Newtons Second Law and Coulombs Law, we see that the acceleration of the alpha particle at a distance r from the nucleus is a= ke (2e)(Ze) F = m m r 2 (1)

in a direction away from the nucleus. (Recall that the alpha particle charge is 2e, where -e is the electrons charge; the nuclear charge is written Ze; m is the alpha-particle mass; and ke is the constant appearing in Coulombs Law.) Hence the scattering angle will depend on ke , Z, e, and m , but only in the single combination2 2ke Ze2 /m (2) These quantities - b, v, and (2) - are the only input parameters on which the scattering angle can depend. Now, is measured in degrees or radians; so its value must be independent of the system of units used to measure distances or times or masses or charges. For instance, we know without doing any calculations at all that the correct formula for is not something like = 1/b or = 1/bv, etc., because numerical values of these quantities do depend on the units used for lengths and times; for instance, if were equal to 1/b, then the scattering angle would be 100 times greater if b were measured in centimeters instead of meters. So the problem is to put together b, v, and (2) in a dimensionless
The distance r between the alpha particle and the nucleus is not included, because it is not one of the input parameters on which might depend, but is rather a dynamical variable, and changes during the collision in a manner governed by Newtons Second Law.
2

combination, that is, in a combination that does not depend on the units used for distances, times, etc. The units of (2) are those of acceleration times distance squared, as can be seen directly from Eq. (1) (just bring the r2 to the left side of the equation). Also, the units of acceleration are distance per time squared (e.g., 9.8 m/sec2 ); so we can also say that the units of (2) are 2ke Ze2 /m (distance)3 /(time)2 (3)

We have no time among our input parameters, but we do have a velocity v whose units are v distance/time (4) To construct a quantity that is independent of the units used to measure time, we must therefore divide (2) by v2 . This yields a quantity with the units 2Zke e2 /m v 2 distance (5) Finally, to construct a quantity that is independent of the units used to measure distance or time, we must divide (5) by the only distance among our inputs, the impact parameter b, and obtain: 2Zke e2 /m v 2 b (6)

Our conclusion is that the scattering angle can depend only on this one combination of input parameters. Equivalently, inverting this relation, we can say that the combination (6) can be expressed as some quantity f() that depends only on the scattering angle: 2Zke e2 /m v 2 b = f () (7)

The impact parameter b() for a given scattering angle is then given by b() = 2Zke e2 /m v 2 f () (8)

Dimensional analysis cannot tell us anything about the nature of the quantity f(), but (8) nevertheless embodies a great deal of information about Rutherford scattering. For instance, if we are interested in scattering by some xed angle , say, 90o , then the impact parameter is doubled if we double the nuclear change Ze, and reduced fourfold if we double the alpha-particle velocity v. Quite a lot to learn with so little work!

Rutherford used Newtonian mechanics to calculate the orbits of the alpha particles scattered by the nucleus, and found that the impact parameter b and deection angle are related by b() =
2Zke e2 m v2 tan(/2)

(9)

This has the general form of (8) that we obtained here by dimensional analysis, and yields the further information that the quantity f() has the value f () = tan(/2) (10)

Here tan is an abbreviation for the angle-dependent quantity known in trigonometry as the tangent: if we draw a right triangle (i.e., one having a 90o or right angle) whose acute angles are and 90o , then tan is the ratio of the side of the triangle opposite the angle whose value is to the side opposite the angle of value 90o . For instance, in a right triangle with both acute angles equal to 45o, the sides opposite these angles are of equal length; so their ratio is one, and therefore tan(45o) = 1. Rutherfords formula (9) then tells us that for = 90o , the impact parameter is b(90o ) = 2Zke e2 m v 2

This is just half the distance of closest approach for an alpha particle red straight at the nucleus, which can be calculated.3 More generally, we may note that (9) gives the impact parameter a plausible dependence on the deection angle. From its interpretation in terms of right triangles, it is apparent that the quantity tan rises steadily from a value 0 at = 0 to innity at = 90o . It follows that b is innite for = 0,
If an alpha particle of charge q = 2e starts at innity with energy E , and has velocity v when it arrives at a distance r from a nucleus with charge q = Ze, then, according to the conservation of energy, the initial energy E must equal the sum of the potential energy V(r) and the kinetic energy 1 mv2 : 2 E = 1 2ke Ze2 + mv 2 r 2
3

For instance, if the alpha particle is directed straight at the nucleus, it will come to rest at a distance rmin given by the solution for v = 0: rmin = 2ke Ze2 E

because zero deection is only possible when the alpha particle has missed the nucleus altogether, drops steadily with increasing , because the closer the collision, the more the deection, and vanishes for = 180o , because the alpha particle must be red straight at the nucleus in order to bounce straight back.

Calculate distribution of deection angle

Suppose that instead of wanting to know the impact parameter for a given deection, or vice versa, we would like to calculate the distribution of deection angles for anpha particles red at random impact parameters into a thin foil. In order to suer a deection greater than a given angle , the alpha particle must have an impact parameter of less than b() for some atomic nucleus in the foil. We can therefore think of b() as the radius of a little disc facing the stream of incoming alpha particles; an alpha particle is deected by more than the angle if it happens to be aimed so that (if it were not deected) it would hit one of these discs. Each disc has an eective area of times its radius squared, or = b()2 (11)

known as the cross section for scattering by at least the angle . To nd out the distribution of deection angles, we must calculate what fraction of the area of the foil is occupied by these discs. The mass M of the foil is equal to the mass m of an individual atom times the number N of atoms in the foil; so N is given by N = M/m (12) Also, the mass of the foil is equal to its density (mass per volumn) times its volume, and the volume of the foil is given by the product of its surface area S and its thickness l; so M = Sl (13) Further, the mass of an individual atom can be expressed as m = A/N0 (14)

where A is the atomic weight and N0 is the quantity known as Avogadros number, dened so that 1/N0 is the mass of a unit atomic weight (1/N0 = 1.671027 kg). Inserting (13) and (14) into (12), we can write the number of atoms in the foil as N = SlN0 /A (15) 9

The probability P() for scattering by more than the angle is given by the fraction of the total area S of the foil that is occupied by the N discs associated with the atoms of the foil, each disc with area (). That is, the scattering probability is P () = N()/S (16) provided the discs do not appreciably overlap. Inserting Eq. (15), we see that the foil area S cancels out, and we have P () = lN0 ()/A (17)

This is a very general formula, applicable to all sorts of scattering processes. For instance, in some (but not all) nuclear reactions, the cross section (0) for scattering by any angle is of the order of the geometric cross-sectional area of the nucleus, or about 21028 kg/m3 and atomic weight 197; so the probability of scattering in gold foil is given by (17) as (2 104 kg/m3 ) l (6 1026 /kg) (2 1028 m2 )/197 = 12l where l (like all the other lenghs here) is expressed in meters. For relatively thick foil with l = 103 m, the scattering probability is 1.2 percent. For thicker foils, the scattering probability approaches unity; that is, the discs begin appreciably to overlap, and the discussion above is no longer applicable. For the special case of Rutherford scattering, the cross section () is given by (9) and (11) as () = 4Z 2 ke 2 e4 /m 2 v 4 [tan(/2)]2 (18)

It follows that the probability (17) of scattering by angle or greater is proportional to 1/[tan /2]2 . Verication of this relation conrms that the force on the alpha particle is really proportional to th einverse square of the distance. (In particular, if the nuclear charge were spread out over a large volume, the cross section and scattering probabilities would vanish more rapidly as approaches 180o.) Also, by using (18) together with (17), we see that the scattering probability is proportional to Z2 ; so measurement of this probability at any given allows one to nd a value for the nuclear charge.

References
[1] Steven Weinberg, The discovery of subatomic particles.

10

You might also like